首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 546 毫秒
1.
Although the role of Bcl‐2 phosphorylation is still under debate, it has been identified in a resistance mechanism to BH3 mimetics, for example ABT‐737 and S1 . We identified an S1 analogue, S1‐16 , as a small‐molecule inhibitor of pBcl‐2. S1‐16 efficiently kills EEE‐Bcl‐2 (a T69E, S70E, and S87E mutant mimicking phosphorylation)‐expressing HL‐60 cells and high endogenously expressing pBcl‐2 cells, by disrupting EEE‐Bcl‐2 or native pBcl‐2 interactions with Bax and Bak, followed by apoptosis. In vitro binding assays showed that S1‐16 binds to the BH3 binding groove of EEE‐Bcl‐2 (Kd=0.38 μM by ITC; IC50=0.16 μM by ELISA), as well as nonphosphorylated Bcl‐2 (npBcl‐2; Kd=0.38 μM ; IC50=0.12 μM ). However, ABT‐737 and S1 had much weaker affinities to EEE‐Bcl‐2 (IC50=1.43 and >10 μM , respectively), compared with npBcl‐2 (IC50=0.011 and 0.74 μM , respectively). The allosteric effect on BH3 binding groove by Bcl‐2 phosphorylation in the loop region was illustrated for the first time.  相似文献   

2.
Antiapoptotic Bcl‐2 family proteins, such as Bcl‐xL, Bcl‐2, and Mcl‐1, are often overexpressed in tumor cells, which contributes to tumor cell resistance to chemotherapies and radiotherapies. Inhibitors of these proteins thus have potential applications in cancer treatment. We discovered, through structure‐based virtual screening, a lead compound with micromolar binding affinity to Mcl‐1 (inhibition constant (Ki)=3 μM ). It contains a phenyltetrazole and a hydrazinecarbothioamide moiety, and it represents a structural scaffold not observed among known Bcl‐2 inhibitors. This work presents the structural optimization of this lead compound. By following the scaffold‐hopping strategy, we have designed and synthesized a total of 82 compounds in three sets. All of the compounds were evaluated in a fluorescence‐polarization binding assay to measure their binding affinities to Bcl‐xL, Bcl‐2, and Mcl‐1. Some of the compounds with a 3‐phenylthiophene‐2‐sulfonamide core moiety showed sub‐micromolar binding affinities to Mcl‐1 (Ki=0.3–0.4 μM ) or Bcl‐2 (Ki≈1 μM ). They also showed obvious cytotoxicity on tumor cells (IC50<10 μM ). Two‐dimensional heteronuclear single quantum coherence NMR spectra of three selected compounds, that is, YCW‐E5, YCW‐E10, and YCW‐E11, indicated that they bind to the BH3‐binding groove on Bcl‐xL in a similar mode to ABT‐737. Several apoptotic assays conducted on HL‐60 cells demonstrated that these compounds are able to induce cell apoptosis through the mitochondrial pathway. We propose that the compounds with the 3‐phenylthiophene‐2‐sulfonamide core moiety are worth further optimization as effective apoptosis inducers with an interesting selectivity towards Mcl‐1 and Bcl‐2.  相似文献   

3.
Turn Bak : We present rationally designed scaffolds that mimic the spatial projection of the i, i+4, i+7, and i+11 residues of an α‐helix. A library of biphenyl derivatives was shown by competition fluorescence polarization and ITC to mimic Bak and disrupt the Bak/Bcl‐xL protein–protein interaction. 15N HSQC experiments confirmed that the surface of Bcl‐xL normally occupied by Bak was the target area of our new synthetic inhibitors.

  相似文献   


4.
In a previous study we reported a class of compounds with a 2H‐thiazolo[3,2‐a]pyrimidine core structure as general inhibitors of anti‐apoptotic Bcl‐2 family proteins. However, the absolute stereochemical configuration of one carbon atom on the core structure remained unsolved, and its potential impact on the binding affinities of compounds in this class was unknown. In this study, we obtained pure R and S enantiomers of four selected compounds by HPLC separation and chiral synthesis. The absolute configurations of these enantiomers were determined by comparing their circular dichroism spectra to that of an appropriate reference compound. In addition, a crystal structure of one selected compound revealed the exocyclic double bond in these compounds to be in the Z configuration. The binding affinities of all four pairs of enantiomers to Bcl‐xL, Bcl‐2, and Mcl‐1 proteins were measured in a fluorescence‐polarization‐based binding assay, yielding inhibition constants (Ki values) ranging from 0.24 to 2.20 μM . Interestingly, our results indicate that most R and S enantiomers exhibit similar binding affinities for the three tested proteins. A binding mode for this compound class was derived by molecular docking and molecular dynamics simulations to provide a reasonable interpretation of this observation.  相似文献   

5.
The design of a cross‐acridine scaffold mimicking the i, i+3, i+5, and i+7 residues distributed over a two‐face, two‐turn α‐helix is described. Docking studies and 2D 1H,15N HSQC NMR spectroscopy provide compelling evidence that compound 3 d accurately reproduces the arrangement of four hotspots in the Bim BH3 peptide to permit binding to the Mcl‐1 and Bcl‐2 proteins (Ki 0.079 and 0.056 μM , respectively). Furthermore, the hotspot mutation could also be mimicked by individual or multiple deletions of side chains on the scaffold.  相似文献   

6.
Zhou B  Li X  Li Y  Xu Y  Zhang Z  Zhou M  Zhang X  Liu Z  Zhou J  Cao C  Yu B  Wang R 《ChemMedChem》2011,6(5):904-921
A class of compounds with a common thiazolo[3,2‐a]pyrimidinone motif has been developed as general inhibitors of Bcl‐2 family proteins. The lead compound was originally identified in a random screening of a small compound library using a fluorescence polarization‐based competitive binding assay. Its binding to the Bcl‐xL protein was further confirmed by 15N‐HSQC NMR experiments. Structural modifications on the lead compound were guided by the outcomes of molecular modeling studies. Among the 42 compounds obtained, a number of them exhibited much improved binding affinities to Bcl‐2 family proteins as compared to the lead compound. The most potent compound, BCL‐LZH‐ 40 , inhibited the binding of BH3 peptides to Bcl‐xL, Bcl‐2, and Mcl‐1 with inhibition constants (Ki) of 17, 534, and 200 nM , respectively.  相似文献   

7.
Considerable efforts have been made to the development of small‐molecule inhibitors of antiapoptotic B‐cell lymphoma 2 (Bcl‐2) family proteins (such as Bcl‐2, Bcl‐xL, and Mcl‐1) as a new class of anticancer therapies. Unlike general inhibitors of the entire family, selective inhibitors of each member protein can hopefully reduce the adverse side effects in chemotherapy treatments of cancers overexpressing different Bcl‐2 family proteins. In this study, we designed four series of benzylpiperazine derivatives as plausible Bcl‐2 inhibitors based on the outcomes of a computational algorithm. A total of 81 compounds were synthesized, and their binding affinities to Bcl‐2, Bcl‐xL, and Mcl‐1 measured. Encouragingly, 22 compounds exhibited binding affinities in the micromolar range (Ki<20 μM ) to at least one target protein. Moreover, some compounds were observed to be highly selective binders to Mcl‐1 with no detectable binding to Bcl‐2 or Bcl‐xL, among which the most potent one has a Ki value of 0.18 μM for Mcl‐1. Binding modes of four selected compounds to Mcl‐1 and Bcl‐xL were derived through molecular docking and molecular dynamics simulations. It seems that the binding affinity and selectivity of these compounds can be reasonably interpreted with these models. Our study demonstrated the possibility for obtaining selective Mcl‐1 inhibitors with relatively simple chemical scaffolds. The active compounds identified by us could be used as lead compounds for developing even more potent selective Mcl‐1 inhibitors with potential pharmaceutical applications.  相似文献   

8.
Managing protein–protein interactions is essential for resolving unknown biological events at the molecular level and developing drugs. We have designed and synthesized a side‐chain‐crosslinked helical peptides based on the binding domain of a pro‐apoptotic protein (Bad) that induces programed cell death. The peptide showed high helical content and bound to its target, Bcl‐XL, more strongly than its non‐crosslinked counterparts. When HeLa cells were incubated with the crosslinked peptide, the peptide entered the cytosol across the plasma membrane. The peptide formed a stable complex with Bcl‐XL localized at the outer mitochondrial membrane, and this binding event caused the release of cytochrome c from the intermembrane space of mitochondria into the cytosol. This activated the caspase cascade: 70 % of HeLa cells died by the apoptosis pathway (without evidence of necrosis).  相似文献   

9.
Targeting Bcl‐x L /Bak : A family of rationally designed α‐helix mimetics with improved solubility and synthetic feasibility based on a benzoylurea scaffold is presented. These benzoylurea derivatives favor a linear conformation stabilized by an intramolecular hydrogen bond, and are able to mimic the spatial projection of the i, i+4, and i+7 residues of an α‐helix. Binding affinities of the benzoylurea derivatives to Bcl‐xL have been assessed using fluorescence polarization competition assays and isothermal titration calorimetry.

  相似文献   


10.
The (R)‐α‐lipoyl‐glycyl‐L ‐prolyl‐L ‐glutamyl dimethyl ester codrug (LA‐GPE, 1 ) was synthesized as a new multifunctional drug candidate with antioxidant and neuroprotective properties for the treatment of neurodegenerative diseases. Physicochemical properties, chemical and enzymatic stabilities were evaluated, along with the capacity of LA‐GPE to penetrate the blood–brain barrier (BBB) according to an in vitro parallel artificial membrane permeability assay for the BBB. We also investigated the potential effectiveness of LA‐GPE against the cytotoxicity induced by 6‐hydroxydopamine (6‐OHDA) and H2O2 on the human neuroblastoma cell line SH‐SY5Y by using the 3‐(4,5‐dimethylthiazol‐2‐yl)‐2,5‐diphenyltetrazolium bromide (MTT) reduction assay. Our results show that codrug 1 is stable at both pH 1.3 and 7.4, exhibits good lipophilicity (log P=1.51) and a pH‐dependent permeability profile. Furthermore, LA‐GPE was demonstrated to be significantly neuroprotective and to act as an antioxidant against H2O2‐ and 6‐OHDA‐induced neurotoxicity in SH‐SY5Y cells.  相似文献   

11.
The kinetics of crystallization induced by orotic acid (OA) and boron nitride (BN) as nucleating agents were investigated for bacterial poly(3‐hydroxybutyrate‐co‐3‐hydroxyhexanoate)s (P(HB‐co‐HH)s) containing from 0 to 18% HH monomer units. The nucleation efficiency of these two chemicals was investigated by differential scanning calorimetry (DSC) and polarized optical microscopy (POM). It was found that both orotic acid and boron nitride are able to nucleate the crystallization of PHB. In the case of P(HB‐co‐HH) copolymers, orotic acid showed an outstanding nucleating effect. The comparison of half‐crystallization times shows that for P(HB‐co‐10% HH), the crystallization initiated by orotic acid is more than three time faster than the one induced by boron nitride (t1/2BN/t1/2OA(60°C) = 3.7 and t1/2BN/t1/2OA(90°C) = 4.5). According to the fact that orotic acid is a biodegradable, biocompatible and a nontoxic chemical, this nucleating agent is a promising solution for PHAs used in medical applications such as implants. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

12.
Refluxing a mixture of phthalonitrile C6R1R2R3R4(CN)2 1 (R1–R4=H), or its substituted derivatives 2 (R1, R3, R4=H, R2=Me), or 3 (R1, R4=H, R2, R3=Cl) (1 equiv.) and N,N‐diethylhydroxylamine, Et2NOH, (4 equivs.) in methanol for 4 h results ( Route A ) in precipitation of the symmetrical ( 6 and 8 ) and an isomeric mixture of unsymmetrical ( 7 ) phthalocyanines, isolated in good (55–65 %) yields. The reaction of phthalonitriles 1 , 2 , or 4 (R1, R3, R4=H, R2=NO2) (4 equivs.) with Et2NOH (8 equivs.) in the presence of a metal salt MCl2 (M=Zn, Cd, Co, Ni) (1 equiv.) in n‐BuOH or without solvent results in the formation of metallated phthalocyanine species ( 9 – 17 ). Upon refluxing in freshly distilled dry chloroform, phthalonitrile 1 or its substituted analogues 2 , 3 or 5 (R1–R4=F) (1 equiv.) react with N,N‐diethylhydroxylamine (2 equivs.) affording 3‐iminoisoindolin‐1‐ones 18 – 21 ( Route B ) isolated in good yields (55–80 %). All the prepared compounds were characterized with C, H, and N elemental analyses, ESI‐MS, IR, and compounds 18 – 21 also by 1D (1H, 13C{1H}), and 2D (1H,15N‐HMBC and 1H,13C‐HMQC, 1H,13C‐HMBC) NMR spectroscopy.  相似文献   

13.
Fucosyltransferases (FucTs) usually catalyze the final step of glycosylation and are critical to many biological processes. High levels of specific FucT activities are often associated with various cancers. Here we report the development of a chemoenzymatic method for synthesizing a library of guanosine diphosphate β‐L ‐fucose (GDP‐Fuc) derivatives, followed by in situ screening for inhibitory activity against bacterial and human α‐1,3‐FucTs. Several compounds incorporating appropriate hydrophobic moieties were identified from the initial screening. These were individually synthesized, purified and characterized in detail for their inhibition kinetics. Compound 5 had a Ki of 29 nM for human FucT‐VI, and is 269 and 11 times more selective than for Helicobacter pylori FucT (Ki=7.8 μM) and for human FucT‐V (Ki=0.31 μM).  相似文献   

14.
Poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) (PHBV) was irradiated by 60Co γ‐rays (doses of 50, 100 and 200 kGy) under vacuum. The thermal analysis of control and irradiated PHBV, under vacuum was carried out by thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). The tensile properties of control and irradiated PHBV were examined by using an Instron tensile testing machine. In the thermal degradation of control and irradiated PHBV, a one‐step weight loss was observed. The derivative thermogravimetric curves of control and irradiated PHBV confirmed only one weight‐loss step change. The onset degradation temperature (To) and the temperature of maximum weight‐loss rate (Tp) of control and irradiated PHBV were in line with the heating rate (°C min?1). To and TP of PHBV decreased with increasing radiation dose at the same heating rate. The DSC results showed that 60Co γ‐radiation significantly affected the thermal properties of PHBV. With increasing radiation dose, the melting temperature (Tm) of PHBV shifted to a lower value, due to the decrease in crystal size. The tensile strength and fracture strain of the irradiated PHBV decreased, hence indicating an increased brittleness. Copyright © 2004 Society of Chemical Industry  相似文献   

15.
BACKGROUND: Poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) [P(3HB‐co‐3HV)] co‐polymer has immense potential in the field of environmental and biomedical sciences as biodegradable and biocompatible material. The present study examines a filamentous N2‐fixing cyanobacterium, Nostoc muscorum Agardh as a potent feedstock for P(3HB‐co‐3HV) co‐polymer production and characterization of co‐polymer film for commercial applications. RESULTS: Under photoautotrophic growth conditions, N. muscorum Agardh accumulated the homopolymer of poly‐β‐hydroxybutyrate (PHB), whereas synthesis of P(3HB‐co‐3HV) co‐polymer was detected under propionate‐ and valerate‐supplemented conditions. Exogenous carbons such as acetate, fructose and glucose supplementation with propionate/valerate was found highly stimulatory for the co‐polymer accumulation; the content reached 58–60% of dry cell weight (dcw) under P‐/N‐deficiencies with 0.4% acetate + 0.4% valerate supplementation, the highest value reported so far for P(3HB‐co‐3HV) co‐polymer‐producing cyanobacterial species. The material properties of the films were studied by mechanical tests, surface analysis and differential scanning calorimetry (DSC). CONCLUSION: N. muscorum Agardh, a photoautotrophic N2‐fixing cyanobacterium, emerged as a potent host for production of P(3HB‐co‐3HV) co‐polymer with polymer content 60% of dry cell weight. The material properties of the films were found to be comparable with that of the commercial polymer, thus advocating its potential applications in various fields. Copyright © 2012 Society of Chemical Industry  相似文献   

16.
2‐Mercapto‐5‐methylpyridine‐N‐oxide (MMPNO) and its sodium salt (NaMMPNO) were synthesized. The reaction of the latter with Fe3+ generates Fe(MMPNO)3 chelate. The thermolysis of this chelate at 350 °C yielded highly pure reddish‐brown γ‐Fe2O3 nanocrystallites with an average particle size of 6.2 nm, a particle size range of 4.2 to 14.8 nm, and a specific surface area of 51.5 m2g–1. The thermolysis process was optimized using the 22 fractional design. Quantitative tests and characterization of products were carried out by UV‐vis spectroscopy, XRD, LLS, SEM, TGA, BET, TEM, FT‐IR, elemental microanalysis, and classical analytical measurements.  相似文献   

17.
4‐(4′‐Aminophenyl)‐1,2,4‐triazolidine‐3,5‐dione ( 1 ) was reacted with 1,8‐naphthalic anhydride ( 2 ) in a mixture of acetic acid and pyridine (3 : 2) under refluxing temperature and gave 4‐(4′‐N‐1,8‐naphthalimidophenyl)‐1,2,4‐triazolidine‐3,5‐dione ( NIPTD ) ( 3 ) in high yield and purity. The compound NIPTD was reacted with excess n‐propylisocyanate in N,N‐dimethylacetamide solution and gave 1‐(n‐propylamidocarbonyl)‐4‐[4′‐(1,8‐naphthalimidophenyl)]‐1,2,4‐triazolidine‐3,5‐dione ( 4 ) and 1,2‐bis(n‐propylamidocarbonyl)‐4‐[4′‐(1,8‐naphthalimidophenyl)]‐1,2,4‐ triazolidine‐3,5‐dione ( 5 ) as model compounds. Solution polycondensation reactions of monomer 3 with hexamethylene diisocyanate ( HMDI ), isophorone diisocyanate ( IPDI ), and tolylene‐2,4‐diisocyanate ( TDI ) were performed under microwave irradiation and conventional solution polymerization techniques in different solvents and in the presence of different catalysts, which led to the formation of novel aliphatic‐aromatic polyureas. The polycondensation proceeded rapidly, compared with conventional solution polycondensation, and was almost completed within 8 min. These novel polyureas have inherent viscosities in a range of 0.06–0.20 dL g?1 in conc. H2SO4 or DMF at 25°C. Some structural characterization and physical properties of these novel polymers are reported. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 90: 2861–2869, 2003  相似文献   

18.
Highly random copolymers of 2,2‐dimethyltrimethylene carbonate (DTC) and ε‐caprolactone (CL) were synthesized by single component rare‐earth tris(4‐tert‐butylphenolate)s [Ln(OTBP)3] for the first time. The influences of reaction conditions on the copolymerization initiated by La(OTBP)3 have been examined in detail. The monomer reactivity ratios of DTC and CL determined by the Fineman–Ross method are 4.0 for rDTC and 0.27 for rCL. The microstructure of the copolymer was determined by the analyses of the diads DTC–DTC, DTC–CL, CL–DTC and CL–CL of the 1H NMR spectra. The high degree of randomness of the chain structure was further confirmed by the 13C NMR spectra and differential scanning calorimetry. The thermal properties of the copolymers as a function of composition are reported. The mechanism investigated by 1H NMR data indicates that the rare‐earth tris(4‐tert‐butylphenolate)s initiate the ring‐opening copolymerization of DTC and CL with acyl‐oxygen bond cleavages of the monomers. Copyright © 2004 Society of Chemical Industry  相似文献   

19.
Herein we report a study aimed at discovering a new class of compounds that are able to inhibit Leishmania donovani cell growth. Evaluation of an in‐house library of compounds in a whole‐cell screening assay highlighted 4‐((1‐(4‐ethylphenyl)‐2‐methyl‐5‐(4‐(methylthio)phenyl)‐1H‐pyrrol‐3‐yl)methyl)thiomorpholine (compound 1 ) as the most active. Enzymatic assays on Leishmania infantum trypanothione reductase (LiTR, belonging to the Leishmania donovani complex) shed light on both the interaction with, and the nature of inhibition by, compound 1 . A molecular modeling approach based on docking studies and on the estimation of the binding free energy aided our rationalization of the biological data. Moreover, X‐ray crystal structure determination of LiTR in complex with compound 1 confirmed all our results: compound 1 binds to the T(SH)2 binding site, lined by hydrophobic residues such as Trp21 and Met113, as well as residues Glu18 and Tyr110. Analysis of the structure of LiTR in complex with trypanothione shows that Glu18 and Tyr110 are also involved in substrate binding, according to a competitive inhibition mechanism.  相似文献   

20.
In this work, we fabricated the (1‐x)BiFeO3xBaTiO3+y‰ mol CuO ceramics by the modified thermal quenching technique. The pure perovskite phase was formed and a morphotropic phase boundary (MPB) was observed in the ceramics with = 0.30‐0.33. The addition of CuO can significantly enhance the density of the BiFeO3‐BaTiO3 material. Importantly, an enhanced piezoelectric constant (d33=165 pC/N), a large electric‐field‐induced strain (?S = 0.54%: peak to peak strain) and a large piezoelectric actuator constant (d33*=449 pm/V) together with a high Curie temperature (TC) of 503°C were observed in the ceramics with = 0.30 and = 5. As a result, the enhanced piezoelectricity and large electric‐field‐induced strain could significantly stimulate further researches in BFO‐based ceramics.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司    京ICP备09084417号-23

京公网安备 11010802026262号